• No se han encontrado resultados

Effects of the precipitation regimen and spatial scale on the invertebrate communities and ecosystem processes in phytotelmata

N/A
N/A
Protected

Academic year: 2020

Share "Effects of the precipitation regimen and spatial scale on the invertebrate communities and ecosystem processes in phytotelmata"

Copied!
199
0
0

Texto completo

(1)

EFFECTS OF THE PRECIPITATION REGIMEN AND SPATIAL SCALE ON THE INVERTEBRATE COMMUNITIES AND ECOSYSTEM PROCESSES IN

PHYTOTELMATA

A Thesis Presented to The Biological Science Faculty

by

Fabiola Ospina Bautista

In partial fulfillment of the requirements for the degree of Doctor of Philosophy in Biological Science

Universidad de los Andes

(2)

2 ABSTRACT

Understanding the factors that drive community structure and ecosystem processes are a

relevant goal in ecology. One factor is environmental heterogeneity that alters

communities through changes in habitat and available resource for species; however, it

is unclear if those effects can change according to the spatial scale. Another factor is

climate change, which affects community composition and ecosystem functionality

through the loss of particular predator species; although, it is unknown if rainfall

variability can alter the community and energy and nutrient flux in temporal

ecosystems, such as ponds or phytotelmata.

We used the community within two phytotelmata, bromeliads and tree holes, as an

ecological model system in order to assess effect of spatial scale and hydrological

regime on communities and ecosystem processes. First, we studied the invertebrate

community associated to Guzmania multiflora (André) André ex Mez. (Bromeliaceae)

and their biological traits. We assessed the relation between biological traits and habitat

complexity and resource availability. We found that habitat complexity not only alters

the taxonomical diversity of invertebrates in bromeliads, but also their functional

diversity through changes in the abundance and richness of biological traits. In this

regard, biological traits provide an approach to ecosystem processes and invertebrate

adaptations to environmental conditions. Second, we investigated the effects of spatial scale dependence of habitat and detritus on community and decomposition. We found

that species turnover of invertebrates associated with tree holes depended on a spatial

scale and that there was a scale-dependent effect of habitat and litter on the community

and litter decomposition. Third, we assessed the relationship between the amount and

(3)

3 the magnitude of precipitation rather than rainfall frequency affected invertebrate

communities, decomposition, and primary productivity. Finally, we analyzed the energy

and nitrogen flux in a bromeliad invertebrate food web and its interaction with the loss

of an intraguild predator. We found that the interaction between shifts in the amount of

precipitation and the presence/absence of the predator altered the energy and nitrogen

(4)

4 TABLE OF CONTENTS

ABSTRACT ... 2

LIST OF TABLES ... 6

LIST OF FIGURES ... 8

ACKNOWLEDGMENTS ... 12

OBJETIVES ... 14

GENERAL OBJETIVES ... 14

SPECIFIC OBJETIVES ... 14

CHAPTER 1. INTRODUCTION ... 15

1.1. REFERENCES ... 21

CHAPTER 2. INVERTEBRATE COMMUNITY ASSOCIATED TO GUZMANIA MULTIFLORA (BROMELIACEA) IN CLOUD MOUNTAIN: HABITAT COMPLEXITY AND AVAILABLE ENERGY AFFECTS BIOLOGICAL TRAITS. ... 35

2.1. ABSTRACT ... 35

2.2. INTRODUCTION ... 35

2.3. METHODS ... 38

2.3.1. Study Area ... 38

2.3.2. Methods ... 38

2.3.3. Statistical analysis ... 39

2.4. RESULTS ... 40

2.5. DISCUSION ... 41

2.5.1. Taxonomical diversity ... 41

2.5.2. Biological traits diversity ... 43

2.6. REFERENCES ... 46

3. CHAPTER 3: SCALE DEPENDENCE OF HABITAT AND LITTER TYPE EFFECTS ON TREE HOLES COMMUNITIES... 67

3.1. ABSTRACT ... 67

3.2. INTRODUCTION ... 67

3.3. METHODS ... 70

3.3.1. Model system ... 70

3.3.2. Experiment 1. Micro scale (Canopy vs. understory)... 71

3.3.3. Experiment 2. Meso scale (Forest vs. plantation) ... 71

3.3.3. Experiment 3. Macro scale (Elevation gradient) ... 72

3.3.4. Data analysis ... 74

3.4. RESULTS ... 75

3.4.1. Micro scale (Canopy vs. Understory) ... 75

3.4.2. Meso scale (Plantation vs Forest) ... 76

3.4.3. Macro scale (Low, Middle, High elevation ... 77

3.5. DISCUSSION ... 78

3.6. REFERENCES ... 80

4. CHAPTER 4: SCALE DEPENDENCE OF HABITAT AND RESOURCE HETEROGENEITY EFFECTS ON DECOMPOSITION IN TREE HOLES ... 91

4.1. ABSTRACT ... 91

(5)

5

4.3. METHODS ... 9594

4.3.1. Micro scale: Canopy vs. understory ... 95

4.3.2. Meso scale: Forest vs. plantation ... 96

4.3.3. Macro scale: Elevation gradient ... 97

4.3.4. Statistical analysis ... 98

4.4. RESULTS ... 98

4.5. DISCUSSION ... 100

4.6. REFERENCES ... 104

CHAPTER 5: PRECIPITATION REGIMEN AFFECTS THE COMMUNITY AND ECOSYSTEM PROCESSES OF BROMELIADS ... 116

5.1. ABSTRACT ... 116

5.2. INTRODUCTION ... 116

5.3. METHODS ... 120

5.3.1. Study area ... 120

5.3.2. Experiment ... 120

5.3.3. Statistical analysis ... 124

5.4. RESULTS ... 126

5.4.1. Environmental conditions ... 126

5.4.2. Invertebrate community ... 126

5.4.3. Bacteria community ... 128

5.4.4. Ecosystem processes ... 129

5.5. DISCUSSION ... 129

5.5.1. Environmental variables ... 130

5.5.2. Invertebrates Community ... 131

5.5.3. Bacteria Community ... 132

5.5.4. Ecosystem processes ... 133

5.6. REFERENCES ... 136

6. CHAPTER 6: HYDROLOGICAL REGIMEN EFFECT ON NITROGEN AND ENERGY FLUX IN A FOOD WEB ... 165

6.1. ABSTRACT ... 165

6.2. INTRODUCTION ... 166

6.3. METHODS ... 169

6.3.1. Study Area ... 169

6.3.2. Statistical analyses ... 172

6.4. RESULTS ... 173

6.4.1. Environmental variables ... 173

6.4.2. Survival of insects ... 173

6.4.3. Flux of energy ... 174

6.4.4. Flux of Nitrogen ... 175

6.5. DISCUSSION ... 175

6.6. REFERENCES ... 180

(6)

6 LIST OF TABLES

Table 2.1. Abundance of invertebrates associated to Guzmania multiflora Habitat: T:

Terrestrial, A: Aquatic; Stage: Ad: Adult, L: Larvae, P: Pupa. ……….………57

Table 2.2. Relation of Leaf number and litter weight with abundance and richness for categories of biological traits of invertebrate into Guzmania multiflora……….61

Table 3.1. Generalized linear models of abundance and number of species associated to

tree hole at different habitat. The variable statistical significant is

highlighted………...87

Table 3.2. Betadiversity Mean and Standard Deviation of the litter treatment in the

elevation gradient experiment ………88

Table 4.1. Effects of response variable on water temperature and pH of tree holes at the

micro, meso, and macro scale. The variable statistical significant is highlighted.

.……….………113

Table 5.1. Effects of mean amount of rainfall per day and dispersion parameter on

(7)

7 Table 5.2. Invertebrate abundance found in the experiment. Habitat T=Terrestrial,

Aq=Aquatic; Stage: A=Adult, I=Immature ……….……….…150

Table 5.3. Generalized linear models for abundance and number of species of macro

invertebrate community with Gaussian family. MU= rainfall per day; K: dispersion

parameter……… 153

Table 5.4. ANOVA Betadiversity Disimilarity Index for abundance and

presence/ausence of macroinvertebrate. MU= rainfall per day; K: dispersion

parameter...……….……… 156

Table 5.5. Effects of mean amount of rainfall per day, dispersion parameter, depth of

water, temperature daily, hydroperiod and days without water on ecosystem process.

GAM; MU= mean amount of rainfall per day; K=dispersion parameter. Highlight

values are significant……….……….………...157

Table 6.1. Generalized linear models assessing the influence of precipitation and trophic

structure treatments on the biomass, abundance of d15N (‰), nitrogen concentration (N

%), and survival for each trophic level in the

(8)

8 LIST OF FIGURES

Figure 2.1. Number of species per invertebrate order found in Guzmania multiflora………..…64

Figure 2.2. Relation between leaf number and abundance (A) and Number of species

(B) of macroinvertebrates associated to Guzmania multiflora ………..…65

Figure 2.3. Abundance and richness of biological traits of macroinvertebrates associated

to Guzmania multiflora A. Abundance of habitat; B. Abundance of stage; C. Abundance

of dispersion; D. Richness Functional groups; E. Abundance of Functional groups.………..………...66

Figure 3.1. Bray Curtis beta diversity according to habitats in each scale. A. Microscale,

B Meso scale, C. Macro scale ……….………89

Figure 3.2. Percentage of variation of community in each spatial scale according to

variable measure in each scale………..………...90

Figure 4.1. Percentage of variation of litter decomposition at each spatial scale

according to habitat, litter, interaction, and not explained ………….………...………114

(9)

9 Figure 4.2. Percentage of remaining litter for each litter and habitat type. A. Micro

scale; B. Meso scale; C. Macro scale Non mix A: Alnus acuminata, Non mix B: Piper

imperiale, Non mix C: Croton magdalenensis, Mix A: A. acuminata and P. imperiale,

Mix B: A. acuminata and C. magdelenensis, Mix C: P. imperiale and C. magdelenensis,

Mix D: A. acuminata, C. magdelenensis, and P.

imperial………..115

Figure 5.1. Precipitation records (mm) of study area for each month from 1997 to 2012 ……..………..159

Figure 5.2. Variation in environmental variable through shift in the amount of rainfall

and dispersion parameter. A. Hydroperiod; B. Days without water; C. Mean of depth of

water; D. Mean of daily temperature………....….160

Figure 5.3. Abundance of invertebrate according to MU (mean rainfall amount) and K

(dispersion parameter). A. Abundace total of invertebrate; B. Abundance of aquatic

invertebrate ………..161

Figure 5.4. Number of invertebrate species according to MU (mean rainfall amount) and

K (dispersion parameter). A. Number of species total; B. Number of species terrestrial;

C. Number of species aquatic………162

Figure 5.5. Shift in the decomposition process according to mean rainfall amount, depth

(10)

10 Variation of FPOM with water depth mean; C. Variation of detritus loss percentage

with hydroperiod. D. Variation of detritus loss percentage with depth mean……….…..163

Figure 5.6. Shift in the Chlorophyll-A and turbidity according to mean rainfall amount

and hydroperiod. A. Variation of Chlorophyll-A with rainfall amount (MU); B.

Variation of turbidity with mean rainfall amount (MU)

………164

Figure 6.1. Trophic structure with species and number of individuals of each trophic

group utilize in the experiment. A. Trophic structure with presence of Oreiallagma

oreas (predator). B. Trophic structure without O. oreas.

……….……….191

Figure 6.2. Shift in biomass of each trophic level into food web according to

precipitation regimen and predator presence. A. Fine particle organic matter mass; B.

Biomass of litter; C. Shift in the filter feeder (Culicidae) biomass; D. Shift in the

detritivores (Scirtes sp.) biomass; E. Shift in the predator (O. oreas) biomass.

Precipitation treatments correspond to 0.1X, 1X and 3x of the rainfall mean. Predator

treatment correspond presence or absence of predator (O. oreas)……….………192

Figure 6.3. Difference in Δδ13C-values (Delta δ13Cfinal−initial) (‰), for each trophic

(11)

11 C (‰) of litter leaf; B. Δδ 13 C(‰) of detritivores (Scirtes sp.); C. Δδ 13 C(‰) of filter

feeder (Culex sp.); D. Δδ 13 C(‰) of predator (O. oreas). Precipitation treatments

correspond to 0.1X, 1X and 3x of the rainfall mean. Predator treatment correspond

presence or absence of predator (O. oreas)

………..………..…....193

Figure 6.4. Shift in the Nitrogen percentage of each trophic level into food web

according to precipitation regimen and predator presence. A. Delta of nitrogen

percentage of litter leaf; B. Nitrogen percentage of Fine particle organic matter mass;

C. Delta of nitrogen percentage of filter feeder (Culex sp.); D. Delta of nitrogen

percentage of detritivores (Scirtes sp.); E. Delta of nitrogen percentage of thepredator

(O. oreas). Precipitation treatments correspond to 0.1X, 1X and 3x of the rainfall mean.

Predator treatment correspond presence or absence of predator (O.

oreas)………...….…194

Figure 6.5. Difference of abundance of Δδ15N (‰), for each trophic level into food web

according to precipitation regimen and predator presence. A. Δδ 15N (‰) of litter leaf;

B. δ15N (‰) of Fine particle organic matter mass; C. Δδ 15N (‰) of filter feeder

(Culex sp.); D. Δδ 15N (‰) of detritivores (Scirtes sp.); E. Δδ15N (‰) of thepredator

(O. oreas). Precipitation treatments correspond to 0.1X, 1X and 3x of the rainfall mean.

Predator treatment correspond presence or absence of predator (O.

(12)

12 ACKNOWLEDGMENTS

First, I wish to express my sincere thanks to my advisor, Emilio Realpe, for giving me

the opportunity to do my PhD within the LAZOEA (Laboratorio de Zoología y Ecología

Acuática). He has given so much of himself to help me succeed. My sincere thanks to

my external advisor Diane Srivastava, professor of Department of Zoology, University

of British Columbia, for many fruitful discussions and ideas for my research; I am

extremely thankful and indebted to her for sharing her ability, and for the sincere and

valuable guidance and encouragement extended to me.

I am grateful to the Department of Biological Sciences of the Universidad de Los Andes

and the Proyecto Semilla Facultad de Ciencias at the Universidad de Los Andes for

providing me with all the necessary facilities for the research. This doctorat would not

have been possible without the economic support of COLCIENCIAS (Doctorado

Nacional 567).

I would also like to express my thanks to Aguas de Manizales for the opportunity to

develop the thesis at the Reserva Forestal Protectora Rio Blanco y Quebrada Olivares.

I am also grateful to University of Utah for giving me the opportunity to take part in

ISOCAMP 2012, and to the ITEC (Inter-university Training for Continental- Scale Ecology) for the fellowship of the project “Effects of drought over nitrogen and energy

flux in a food web”. I would also like to express my sense of gratitude to Jed P. Sparks, Director of the Cornell Isotope Laboratory (COIL) and professor of Ecology or

Evolutionary Biology at the University of Cornell, and to Kimberlee Sparks, Cornell

(13)

13 analysis of stable isotope and discussion about stable isotope.

I also thank Kurtis Trzcinski, postdoc student at the University British Columbia, who

helped in the field and for his suggestion in the precipitation regime research. I take this

opportunity to express my gratitude to all field assistants for their help and support in

the field station.

(14)

14 OBJETIVES

GENERAL OBJETIVES

Assess the effects of the hydrological regimen and spatial scale on invertebrate

communities and ecosystem processes of phytotelmata.

SPECIFIC OBJETIVES

 Characterize the invertebrate community associated to Guzmania

multiflora (Bromeliaceae) in Cloud Mountain.

 Assess the scale dependence of habitat and litter type effects on tree

holes communities.

 Assess the scale dependence of habitat and litter type effects on

decomposition processes.

 Determine the changes in community and ecosystem processes produced

by a change in the frequency and magnitude of rainfall.

 Assess the effects of rainfall magnitude and trophic structure on energy

(15)

15 CHAPTER 1. INTRODUCTION

Phytotelmata are small aquatic habitats formed by different plant structures such as leaf

axils, fruits, flower parts, or tree trunks and branches (Fish, 1983; Kitching, 2000).

Examples of phytotelmata include pitcher plants, bamboo internodes, tree holes, and

tank bromeliads (Kitching, 2001). Bromeliads and tree holes are the most abundant

phytotelmata in the Neotropics; the rosette disposition of bromeliad leaves creates a

tank that allows accumulation of rainwater and detritus from the canopy. The rainwater

reserve, however, is not constant throughout the bromeliad, thus, the bromeliad provides

two habitats for fauna: the young leaf axils maintain water in their interior forming an

aquatic habitat, whereas water is lost by older or mature leaves in the outer part of

bromeliad provides a terrestrial habitat (Araujo et al., 2007; Montes de Oca et al., 2007;

Frank, 1983). Tree holes are formed by the rainwater catchment of decomposing

cavities or depressions in the woody portions of trees (Kitching, 1971). Tree holes

receive organic material from trapped dead leaves and inorganic ions from stem flow

(Eaton et al., 1973).

Overall, phytotelmata contain small volumes of water and detritus that allow the

maintenance of associated communities (Fish, 1983; Maguire, 1970; Srivastava et al.,

2004). The chemical features of water are influenced by decomposing plant material

and the metabolic activity of organisms (Ngai and Srivastava, 2006); therefore, water in

phytotelmata has low oxygen concentration and an acidic pH (Laessle, 1961).

Phytotelmata are characterized by a defined area as a discrete and replicable unit

(16)

16 community on a temporal and spatial scale (Kitching, 1971). They contain fauna with

short life cycles (Srivastava et al., 2004) and, in some cases, only a part of the life cycle

is developed in the phytotelmata, allowing species dispersion and energy flux between

phytotelmata and terrestrial ecosystems (Giller et al., 2004). Phytotelmata have lower

species richness than macroecosystems, allowing a clearly defined food web. These

characteristics have led to an increase in phytotelmata studies focused on ecosystem,

community, and population ecology, showing the relevance of these microecosystems

in ecology.

The first studies on bromeliads and tree holes focused on species reports, especially of

taxonomic groups such as damselflies and mosquitoes (Frank, 1983; Galindo et al.,

1950, 1951, 1955; Kitching, 1971;Laessle, 1961; Lounibos et al., 1987; Picado, 1913;

Pittendrigh, 1948 ). These studies and others have reported that communities in

bromeliads and tree holes are characterized by bacteria (Brighigna et al., 1992 ;Goffredi

et al., 2011a,b; Vega-Sepulveda, 2009), algae, flagellates, ciliates (Carrias, 2001;

Durán-Ramírez et al., 2015), arthropods, and vertebrates (Neill, 1951). Arthropods

show high diversity and abundance in bromeliads and tree holes (Kitching, 2000; Liria,

2007; Ospina-Bautista et al., 2004; Ospina-Bautista et al., 2008; Richardson et al.,

2000; Williams, 2006; Yanoviak et al., 2006a), with the orders Diptera (Cranstron,

2007; Derraik, 2005; Epler and Janetzky, 1998; Frank and Lounibos, 2009; Wagner et

al., 2008), Odonata (Corbet, 1983; De Marmels and Garrison, 2005; Melnychuk and

Srivastava, 2002), Hemiptea (Polhemus and Polhemus, 1991), Coleoptera (Blake et al.,

2008; Ospina–Bautista et al., 2004), Collembola, Orthoptera (Frank et al., 2004),

Blattodea, and Dermaptera (Frank and Lounibos, 2009). Moreover, diplopods,

(17)

17 2005), mites (Frank et al., 2004), scorpions, and pseudo scorpions (Frank et al., 2004;

Richardson, 1999) have been reported.

These organisms are important for nutrient cycling and ecosystem processes. Terrestrial

fauna that visit the bromeliad, such as ants, spiders, and frogs, may contribute to

nitrogen cycling by depositing nitrogen to the bromeliad through their feces (Leroy et

al., 2009; Romero et al., 2010; Romero and Srivastava, 2010). Moreover, predation

activity of damselflies (Mecistogaster modesta) retains the nitrogen that can be lost

through detritivorous emergence on bromeliads; therefore, damselflies are an important

element of the nitrogen flux from leaf litter to the bromeliad (Ngai and Srivastava,

2006). Predators in bromeliads and their interactions may affect the carbon cycle

through their influence on carbon dioxide concentrations in bromeliads; for example,

damselflies reduce carbon dioxide concentration, but their interaction with another

predator (Copelatus sp.) eliminates top-down influences on the carbon cycle (Atwood et

al., 2014). In addition, bacteria such as Methanomicrobiales, Methanocelales, and

Metanosarcinales release methane as a result of decomposition processes, turning the

bromeliad into a methane source (Goffredi et al., 2011b).

Bromeliad and tree holes communities have been used to assess the abiotic and biotic

factors that determine communities and populations. Overall, researchers have found

that habitat characteristics are the most relevant factors determining populations and

community structure. For instance, habitat size determines spider assemblages,

protozoa, algal, and invertebrate communities (Goncalves-Souza et al., 2011; Carrias et

al., 2001; Carrias et al., 2014; Frank, 2004;), as well as the population density of

(18)

18 holes (Yanoviak, 1999), and Culex biscaynensis in exotic bromeliads (O’Meara et al.,

2003). Tree hole depth and height determine mosquito distribution (Copeland and

Craig, 1990) and relevant characteristics of oviposition selection (Sinsko and

Grimstand, 1977). In addition, habitat complexity, measured through leaf number,

determines algal, protozoa, and invertebrate communities, showing a positive relation

with protozoa richness and invertebrate communities, and a negative relation with algal

communities (Armbruster et al., 2002; Carrias, 2001; Carrias et al, 2012; Jabiol et al.,

2009). Moreover, bromeliad leaf architectures affect the presence of Psecas chapoda

(Salticidae) (O’mena and Romero, 2008). Finally, habitat location influences the

abundance and richness of invertebrates associated with tree holes and is related to

habitat selection by adults (Fincke and Yanoviak, 1997).

Likewise, water volume and its physicochemical features can affect populations or

communities. Water volume in phytotelma has a positive relation with invertebrate and

algal richness and abundance (Araujo, 2007; Carrias et al., 2014; Dézerald et al., 2014;

Jabiol et al., 2009; Kitching, 2001;Yanoviak et al., 2006a), as well as Chironomidae

abundance (Sodré et al., 2010) and Culicidae assemblages (Marques et al., 2012).

Variation in pH appears to be a relevant condition for bacteria and insects; in this

regard, Alphaproteobacteria, Acidobacteria, Planctomycetes, Bacteroidetes,

Betaproteobacteria, Firmicutes, and Bacteroidetes of bromeliads in Costa Rica show a

positive or negative relation to water pH according to the bacterial species (Goffredi et

al., 2011a); for example, a low pH leads to slow growth of Helodes pulchella (Paradise,

(19)

19 Detritus is the main base resource in tree holes and bromeliads, and strongly influences

freshwater faunal communities (Dézerald et al., 2014). Studies have found that low

quality and quantity of this resource leads to lower invertebrate richness and abundance

(Fish and Carpenter, 1982; Richardson, 1999). However, Srivastava and Lawton (1998)

found that an increase in detritus leads to more species, but not greater abundance in

tree holes. Detrital variations shift oviposition of mosquitoes such as Aedes albopictus

and Aedes aegypti (Fader and Juliano, 2014). Detrital quantity alters the growth and

survival of Helodes pulchella, which shows slower growth when fed less detritus, but

reduced survival under high detrital amounts in tree holes (Paradise, 1999).

Chironomidae and Tricoptera growth in bromeliads is limited by the nutrients in leaf

litter (Gonzalez et al., 2014). Moreover, high values of fine particule organic matter

(FPOM) lead to an increase protozoan richness (Carrias et al., 2012).

Bromeliads and tree holes communities have been used to assess the effects of

predation, herbivory and facilitation on species and communities. Aquatic predators

reduce the abundance and richness of invertebrates (Yanoviak, 2001a), rotifers, and

protozoa (Kneitel and Chase, 2011) in tree holes, whereas spiders regulate abundance

and diversity of invertebrates that carry out their entire life cycle in bromeliad water

(Romero and Srivastava, 2010). Predation also affects population dynamics; for

instance, a top predator (Toxorhynchites rutilus) reduces survivorship of prey species

and intermediate predators (Corethrella appendiculata), changing prey composition in

tree holes (Griswold and Lounibos, 2006). Mecistogaster modesta (top predator) shows

a similar effect on Chironomidae emergence (collector-gatherer) (Starzomski et al.,

2010) and mosquitos in bromeliads (Hammill et al., 2015), while the feeding activity of

(20)

20 by the bromeliad-eating weevil (Metamasius callizona) decreases the abundance of

Tillandsia utriculata in the canopy, reducing available aquatic habitats for communities

(Cooper et al., 2014).

Facilitation processes have been reported for bromeliads and tree holes. In bromeliads,

detritivory facilitates the emergence of Chironomidae (Starzomski et al., 2010);

furthermore, ciliate communities change with the loss of detritivore diversity when the

top predator is extinct (Srivastava and Bell, 2009). The presence of ants (Camponotus

femoratus) produces greater richness and abundance of protists (Carrias et al., 2012),

and frogs and snakes help to disperse ostracods (Serramo et al., 1999). In tree holes,

feeding by Scirtidae beetles indirectly facilitates mosquito production through the

microbial community (Daugherty and Juliano, 2003; Pelz-Stelinski et al., 2011).

Climatic conditions and anthropogenic effects lead to habitat loss and altered water and

nutrient cycles and availability of bromeliads and tree holes. Climatic conditions such as

hurricanes decrease the alpha and beta diversity of associated invertebrates, since

hurricanes reduce bromeliad density and rare species (Richardson et al., 2015). Drought

has the greatest effects on diversity and abundance of tree holes insects within forest

patches and bromeliads (Amundrud and Srivastava, 2015; Srivastava, 2005), while

anthropogenic effects such as deforestation, land use, and pollution produce shifts in

communities. For instance, deforestation increases the abundance of mosquitoes in tree

holes (Yanoviak et al., 2006b), land use leads to different community composition

(Yanoviak et al., 2006b); and pollutant as pentachlorophenol alters the bacterial

(21)

21 In summary, phytotelmata are micro-aquatic ecosystems in terrestrial ecosystems with

special features that allow them to harbour an important diversity of bacteria, protists,

invertebrates, and vertebrates that support ecosystem processes; therefore, phytotelmata

could be used as an ecological model (Srivastava et al., 2004). As phytotelmata,

bromeliads and tree holes could be relevant to assess the effect of the hydrological

regimen and spatial scale on phytotelmata communities.

1.1. REFERENCES

Ager, D., Evans, S., Li, H., Lilley, A.K. and Van der Gast, C.J. 2010. Anthropogenic

disturbance affects the structure of bacterial communities. Environmental Microbiology

12:670-678.

Araujo, V., Melo, S., Araujo, A., Gomes, M. and Carneiro, M. 2007. Relationship

between invertebrate fauna and bromeliad size. Brazilian Journal of Biology 67:611–

617.

Amundrud, S.L., Srivastava, D.S., and O'Connor, M.I. 2015. Indirect effects of

predators control herbivore richness and abundance in a benthic eelgrass (Zostera

marina) mesograzer community. Journal of Animal Ecology 84(4): 1092-1102.

Armbruster, P., Hutchinson R.A. and Cotgreave, P. 2002. Factors influencing

(22)

22 Atwood, T.B., Hammill, E., Srivastava, D.S. and Richardson, J.S. 2014.Competitive

displacement alters top-down effects on carbon dioxide concentrations in a freshwater

ecosystem. Acta Oecologia 175:353–361.

Blake, M., Gómez-Zurita, J., Ribera, I., Vitoria, A., Zillikens, A., Steiner, J., García, M.,

Hendrich, L. and Vogler, A.P. 2008. Ancient associations of aquatic beetles and tank

bromeliads in the Neotropical forest canopy. Proceedings of the National Academy of

Sciences 105:6356–6381.

Brighigna, L., Montaini, P., Favilli, F. and Trejo, A.C. 1992. Role of the

nitrogen-fhxing bacterial microflora in the epiphytism of Tillandsia (Bromeliaceae). Journal of

Botany 79:723–727.

Carrias, J.F., Cereghino, R., Brouard, O., Pelozuelo, L., Dejean, A., Coute, A., Corbara,

B. and Leroy, C. 2014. Two coexisting tank bromeliads host distinct algal communities

on a tropical inselberg. Plant Biology 16:997–1004.

Carrias, J.F., Brouarda, O., Leroy, C., Céréghino, R., Pélozuelod, L., Dejean, A.,

Corbara, B. 2012. An ant–plant mutualism induces shifts in the protist community

structure of a tank-bromeliad. Basic and Applied Ecology 13:698–705.

Carrias, J., Cussac, M.E. and Corbara, B. 2001. A preliminary study of freshwater

(23)

23 Cooper, T.M., Frank, J.H. and Cave, R.D. 2014. Loss of phytotelmata due to an

invasive bromeliad-eating weevil and its potential effects on faunal diversity and

biogeochemical cycles. Acta Oecologica 54:51–56.

Copeland, R.S. and Craig, G.B. Jr. 1990. Habitat segregation among treehole

mosquitoes (Diptera: Culicidae) in the Great Lakes region of the United States. Annals

of the Entomological Society of America 83:1O63-1O73.

Corbet, P.S. 1983. Odonata in phytotelmata. En: Frank, J.H. and Lounibos, L.P., (eds.)

Phytotelmata: terrestrial plants as hosts for aquatic insect communities. Plexus

publishing, New Jersey. 29-54p.

Cranston, P.S. 2007. New Species for a Bromeliad Phytotelm-Dwelling Tanytarsus (Diptera : Chironomidae). Annals of the Entomological Society of America 100:617–

622.

Daugherty, M.P. and Juliano, S.A. 2003. Leaf-scraping beetle feces are a food resource

for treehole mosquito larvae. American Midland Naturalist 150:181–184.

De Marmels, J. and Garrison, R.W. 2005. Review of the genus Leptagrion in Venezuela

with new synonymies and descriptions of a new genus, Bromeliagrion, and a new

species B. rehni (Zygoptera: Coenagrionidae). The Canadian Entomologist 137:257–

(24)

24 Derraik, J.G.B. 2005. Mosquitoes breeding in phytotelmata in native forests in the

Wellington region, New Zealand. New Zealand Journal Ecology 29:185–191.

Dézerald, O., Talaga, S., Leroy, C., Carrias, J.F., Corbara, B., Dejean, A., Cereghino, R.

2014. Environmental determinants of macroinvertebrate diversity in small water bodies:

Insights from tank-bromeliads. Hidrobiologica 723:77–86.

Durán-Ramírez, C.A., García-Franco, J.G., Foissner, W. and Mayén-Estrada, R. 2015.

Free-living ciliates from epiphytic tank bromeliads in Mexico. European Journal of

Protistology 51:15–33.

Eaton, J.S., Likens, G.E. and Bormann, F.H. 1973. Throughfall and stem-flow

chemistry in a northern hard- wood forest. Journal of Ecology61:495-508.

Epler, J.H. and Janetzky, W.J. 1998. A new species of Monopelopia (Diptera:

Chironomidae) from phytotelmata in Jamaica, with preliminary ecological notes.

Journal of the Kansas Entomological Society 71:216–225.

Fader, J.E. and Juliano, S.A. 2014. Oviposition habitat selection by container-dwelling

mosquitoes: Responses to cues of larval and detritus abundances in the field. Ecological

Entomology 39:245–252.

Fincke, O., Yanoviak, S.P. 1997. Predation by odonates depresses mosquito abundance

(25)

25 Fish, D. 1983. Phytotelmata: Flora and Fauna. En: Frank, J.H. and Lounibos, L.P.,

(eds.) Phytotelmata: terrestrial plants as hosts for aquatic insect communities. Plexus

publishing, New Jersey. 1-25p.

Fish, D. and Carpenter, S.R. 1982. Leaf litter and larval mosquito dynamics in tree-Hole

ecosystems. Ecology 63:283–288.

Frank, J.H. and Lounibos, L.P. 2009. Insects and allies associated with bromeliads: a

review. Terrestrial Arthropod Reviews 1: 125–153.

Frank, J.H., Reenivasan, S.S., Benshoff, P.J., Deyrup, M.A., Edwards, G.B., Halbert,

S.E., Hamon, A.B., Lowman, M.D., Mockford, E.L., Scheffrahn, R.H., Steck, G.J.,

Thomas, M.C., Walker, T.J. and Welbourn, W.C. 2004. Invertebrate animals extracted from native tillandsia (Bromeliales : Bromeliaceae) in Sarasota County, Florida. Florida

Entomological 87:176–185.

Frank, J.H. 1983. Bromeliad phytotelmata and their biota, especially mosquitos,

pp.101-128. En: Frank, J.H., Lounibos, L.P. (eds) Phytotelmata: terrestrial plants as hosts of

aquatic insects communities. Plexus publishing, New Jersey. 293p.

Galindo P., Carpenter, S.J. and Trapido, H. 1955. A contribution to the ecology and

biology of tree hole breeding mosquitoes of Panama. Annals of the Entomological

(26)

26 Galindo, P., Carpenter, S.J. and Trapido, H. 1951. Ecological observations on forest

mosquitoes of an endemic yellow fever area in Panama. The American Journal of

Tropical Medicine and Hygiene 31:98-137.

Galindo, P., Carpenter, S.J. and Trapido, H. 1950. Observations on diurnal forest

mosquitoes in relation to sylvan yellow fever in Panama. The American Journal of

Tropical Medicine and Hygiene 30:533-574.

Giller, P.S., Hillebrand, H., Berninger, U.G., Gessner, M.O., Hawkins, S., Inchausti, P.,

Inglis, C., Leslie, H., Malmqvist, B., Monaghan, M.T., Morin, P.J. y O’Mullan, G.

2004. Biodiversity effects on ecosystem functioning: emerging issues and their

experimental test in aquatic environments. Oikos 104:423-436.

Goffredi, S.K., Kantor, A.H., and Woodside, W.T. 2011a. Aquatic microbial habitats

within a neotropical rainforest: bromeliads and pH-associated trends in bacterial

diversity and composition. Microbial Ecology 61:529-542.

Goffredi, S.K., Jang, G.E., Woodside, W.T. and Ussler, W. 2011b. Bromeliad

catchments as habitats for methanogenesis in tropical rainforest canopies. Frontiers in

Microbiology 2:256.

Goncalves-Souza, T., Almeida-Neto, M. and Romero, G.Q. 2011. Bromeliad

architectural complexity and vertical distribution predict spider abundance and richness.

(27)

27 Gonzalez, A.L., Romero, G. and Sirivastava, D.S. 2014. Detrital nutrient content

determines growth rate and elemental composition of bromeliad-dwelling insects.

Freshwater Biology 59:737–747.

Griswold, M.W. and Lounibos, L.P. 2006. Predator identity and additive effects in a

treehole community. Ecology 87:987–995.

Hammill, E.D.D., Atwood, T.B., Corvalan, P. and Srivastava, D.S. 2015. Behavioural

responses to predation may explain shifts in community structure. Freshwater Biology

60:125–135.

Jabiol, J., Corbara, B., Dejean, A. and Céréghino, R. 2009. Structure of aquatic insect

communities in tank-bromeliads in a East-Amazonian rainforest in French Guiana.

Forest Ecology and Management 257:351–360.

Kitching, R.L. 2001. Food webs in phytotelmata: “Bottom-Up” and “Top-Down”

explanations for community structure. Annual Review of Entomology 46:729-760.

Kitching, R.L., 2000. Food webs and container habitats: the natural history and ecology

of Phytotelmata. Cambridge University Press, Cambridge. 431p.

Kitching, R.L. 1971. An ecological study of water-filled tree-holes and their position in

(28)

28 Kneitel, J.M. and Chase, J.M. 2011. Disturbance, predator, and resource interactions

alter container community composition. Ecology 85:2088–2093.

Laessle, A.M. 1961. A Micro-Limnological Study of Jamaican Bromeliads. Ecology

42:499–517.

Leroy, C., Corbara, B., Dejean, A. and Cereghino, R. 2009. Ants mediate foliar

structure and nitrogen acquisition in a tank- bromeliad. New Phytologist 183:1124–

1133.

Liria, J. 2007. Fauna fitotelmata en las bromelias Aechmea fendleri André y

Hohenbergia stellata Schult del Parque Nacional San Esteban, Venezuela. Revista

Peruana de Biología 14:33–38.

Lounibos, L.P., Frank, J.H., Machado–Allison, C.E., Navarro, J.C. and Ocanto, P. 1987.

Seasonality, abundance and invertebrate associates of Leptagrion siqueirai Santos

in Aechmea bromeliads in Venezuelan rain forest (Zygoptera:

Coenagrionidae). Odonatologica 16:193–199.

Maguire, B. 1970. Aquatic communities in bromeliad leaf axils and the influence of

radiation. Pp. E.95–E101. En: Odum, H.T. and Pigeon, R.F. (eds). A tropical rain forest.

A study of irradiation and ecology at El Verde, Puerto Rico. Division of Technical

(29)

29 Marques, T.C., Bourke, B.P., Laporta, G.Z. and Sallum, M.A.M. 2012. Mosquito

(Diptera: Culicidae) assemblages associated with Nidularium and Vriesea bromeliads in

Serra do Mar, Atlantic Forest, Brazil. Parasites and Vectors 5:41.

Melnychuk, M.C and Srivastava, D.S. 2002. Abundance and vertical distribution of a

bromeliad-dwelling zygopteran larva, Mecistogaster modesta, in a Costa Rican

rainforest (Odonata: Pseudostigmatidae). International Journal of Odonatology 5:81–97.

Montes de Oca, E., Ball, G.E. and Spence, J.R. 2007. Diversity of Carabidae (Insecta,

Coleoptera) in epiphytic Bromeliaceae in Central Veracruz, Mexico. Environmental

Entomology 36:560–568.

Neill, W.T. 1951. A Bromeliad herpetofauna in Florida. Ecological Society of America

32:140–143.

Ngai, J.T. and Srivastava, D.S. 2006. Predators accelerate nutrient cycling in a

Bromeliad Ecosystem. Science 314:963.

O’Meara, G.F.O., Cutwa, M.M. and Evans, L.F. 2003. Bromeliad-inhabiting

mosquitoes in south Florida: native and exotic plants differ in species composition.

Journal of Vector Ecology 28:37–46.

O’mena, P.M. and Romero G.Q. 2008. Fine-scale microhabitat selection in a

bromeliad-dwelling jumping spider (Salticidae). Biological Journal of the Linnean Society 94:653–

(30)

30 Ospina-Bautista, M.F., Estevez-Varón, J.V., Betancur, J. and Realpe, E. 2004.

Macroinvertebrados acuáticos asociados a la bromelia Tillandsia turneri en un bosque

altoandino. Revista Acta Zoológica Mexicana 20(1):153-166.

Ospina-Bautista, M.F., Estevez-Varon, J.V., Realpe, E. and Gast, F. 2008. Diversidad

de invertebrados acuáticos asociados a Bromeliaceae en un bosque de montaña. Revista

Colombiana de Entomología 34(2):224-229.

Paradise, C.J. 2000. Effects of pH and resources on a processing chain in simulated

interaction treeholes. British Ecological Society 69:651–658.

Paradise, C.J. and Kuhn, K.L. 1999. Interactive effects of pH and leaf litter on a

shredder, the scirtid beetle, Helodes pulchella, inhabiting tree-holes. Freshwater

Biology 41:43–49.

Pelz-Stelinski, K.S., Kaufman, M.G. and Walker, E.D. 2011. Beetle (Coleoptera: Scirtidae) facilitation of larval mosquito growth in tree hole habitats is linked to

multitrophic microbial interactions. Microbial Ecology 62:690-703.

Picado, C. 1913. Les broméliacées épiphytes considerées comme milieu biologique.

Bulletin des Sciences de la France et de la Belgique 47:215–360.

Pittendrigh, C.S. 1948. The bromeliad-Anopheles-malaria complex in Trinidad; The

(31)

31 Polhemus, J.T. and Polhemus, D.A. 1991. A review of the veliid fauna of bromeliads

with a key and description of a new species (Heteroptera Veliidae). Journal of the New

York Entomological Society 99:204–216.

Richardson, M.J., Richardson, B.A. and Srivastava, D.S. 2015. The Stability of

invertebrate communities in bromeliad phytotelmata in a rain forest subject to

hurricanes. Biotropica 47:201–207.

Richardson, B.A., Rogers, C. and Richardson, M.J. 2000. Nutrients, diversity, and

community structure of two phytotelm systems in a lower montane forest, Puerto Rico.

Ecological Entomology 25:348–356.

Richardson, B.A. 1999. The Bromeliad microcosm and the assessment of faunal

diversity in a neotropical forest.Biotropica 31(2): 321-336.

Romero, G.Q. and Srivastava, D.S. 2010. Food-web composition affects

cross-ecosystem interactions and subsidies. Journal of Animal Ecology 79:1122–1131.

Romero, G.Q., Nomura, F., Gonçalves, A.Z., Dias, N., Mercier, H., Conforto, E. de C.,

Rossa-Feres, D. de C. 2010. Nitrogen fluxes from treefrogs to tank epiphytic

bromeliads: An isotopic and physiological approach. Acta Oecologia 162:941–949.

Romero, G.Q. and Vasconcellos-Neto, J. 2005. Spatial distribution and microhabitat

(32)

32 Arachnology 33:124–134.

Serramo-Lopez, L.C., Pena-Rodrigues, P.J.F. and Iglesias-Rios, R. 1999. Frogs and

snakes as phoretic dispersal agents of bromeliad ostracods (Limnocytheridae: Elpjdium)

and annelids (Naididae: Dero). Biotropica 31:705–708.

Sinsko, M.J. and Grimstad, P.R. 1977. Habitat separation by differential vertical

oviposition of two treehole Aedes in Indiana. Environmental Entomology 6:485-487.

Sodré, V.M., Rocha, O. and Messias, M.C. 2010. Chironomid larvae inhabiting

bromeliad phytotelmata in a fragment of the Atlantic Rainforest in Rio de Janeiro State.

Brazilian Journal of Biology 70(3):587-92.

Srivastava, D.S. and Bell, T. 2009. Reducing horizontal and vertical diversity in a food

web triggers extinctions and impacts functions. Ecology Letters 12:1016–1028.

Srivastava, D.S. 2005. Do local processes scale to global patterns? The role of drought

and the species pool in determining treehole insect diversity. Acta Oecologia 145:205–

215.

Srivastava D.S, Kolasa, J., Bengtsson, J., Gonzalez, A., Lawler, S.P., Miller, T.E.,

Munguia, P., Romanuk, T., Schneider, D.C. and Trzcinski, M.K. 2004. Are natural

microcosms useful model systems for ecology?.Trends in Ecology and Evolution 19(7):

(33)

33 Srivastava, D.S. and Lawton, J.H. 1998. Why more productive sites have more Species :

An experimental test of theory using tree-hole communities. The American Naturalist

152:510–529.

Starzomski, B.M., Suen, D. and Srivastava, D.S. 2010. Predation and facilitation

determine chironomid emergence in a bromeliad-insect food web. Ecological

Entomology 35:53–60.

Vega-Sepulveda, J.A. 2009. Bacterias fototróficas anoxigénicas púrpuras no-sulfurosas

en la fitotelmata de bromelias en diversos bosques de Puerto Rico. Tesis Maestría,

Universidad de Puerto Rico, Recinto Universitario de Mayagüez. Puerto Rico. 104 pp.

Wagner, R., Richardson, B.A. and Richardson, M.J. 2008. A new psychodid species

from Puerto Rican tank bromeliads. Studies on Neotropical Fauna and Environment

43:209–216.

Walker, E.D., Lawson, D.L., Merritt, R.W., Morgan, W.T. and Klug, M.J.

1991. Nutrient dynamics, bacterial-populations, and mosquito productivity in tree hole

ecosystems and microcosms. Ecology 72:1529–1546.

Williams, D. 2006. The Biology of Temporary Waters. University of Toronto at

(34)

34 Yanoviak, S.P., Lounibos, L.P. and Weaver, S.C. 2006a. Land use affects

macroinvertebrate community composition in phytotelmata in the peruvian amazon.

Annals of the Entomological Society of America 99(6):1172–1181.

Yanoviak, S.P., Ramírez-Paredes, J.E., Lounibos, L.P. and Weaver, S.C. 2006b.

Deforestation alters phytotelm habitat availability and mosquito production in the

peruvian amazon. Ecogical Applications 16:1854–1864.

Yanoviak, S.P. 2001.. Predation, resource availability, and community structure in

Neotropical water-filled tree holes. Acta Oecologia 126:125–133.

Yanoviak, S.P. 1999. Distribution and abundance of Microvelia cavicola Polhemus

(Heteroptera: Veliidae) on Barro Colorado Island, Panama. Journal of the New York

(35)

35 CHAPTER 2. INVERTEBRATE COMMUNITY ASSOCIATED TO GUZMANIA

MULTIFLORA (BROMELIACEA) IN CLOUD MOUNTAIN: HABITAT

COMPLEXITY AND AVAILABLE ENERGY AFFECTS BIOLOGICAL TRAITS.

2.1. ABSTRACT

Habitat complexity and available energy are among the most important factors

structuring communities. Moreover, these factors could determine the biological traits

of organisms in a community, thus, altering ecosystem processes. We studied the

relation between habitat complexity and resource availability and the biological traits of

the invertebrate community in Guzmania multiflora (André) André ex Mez.

(Bromeliaceae), using leaf number and the amount of litter contained as measures of

habitat complexity and available energy, respectively. We collected the inhabiting

invertebrates and determined their biological traits (habitat, stage, functional group, and

dispersal type). We found that habitat complexity not only alters the taxonomical

diversity of invertebrates in bromeliads, but also their functional diversity through

changes in the abundance and richness of biological traits. In conclusion, biological

traits provide an approach to ecosystem processes and invertebrate adaptations to

environmental conditions; therefore, the study of changes in the abundance and richness

of biological traits could contribute key tools to study the effects of climate change and

anthropogenic disturbances on ecosystems.

2.2. INTRODUCTION

Habitat complexity and available energy are among the most important factors

structuring communities (Hurlbert, 2004). Habitat complexity alters the richness and

(36)

36 more complex habitats (Kovalenko et al., 2012). This tendency has been explained

through the niche space hypothesis, which states that complex habitats give more niche

space to species (Willis et al., 2005), and through the resilience hypothesis where

habitat complexity could lead to an increased efficiency in utilizing resources and

providing greater resilience to communities from disturbances (Kovalenko et al., 2012).

On the other hand, energy availability could determine the number of species in

communities (Species-energy theory, Wright, 1983; Tittensor et al., 2011) by increasing

the number of individuals in the community, reducing the stochastic risk of species

extinction, and supporting more species (Srivastava and Lawton, 1998; Hurlbert, 2006),

or by a high resource abundance that can promote the occurrence of specialist species

increasing richness and reducing competition (Evans et al., 2005).

Moreover, habitat complexity and resource availability can alter ecosystem processes.

The effect of habitat complexity on ecosystem processes is poorly known (Tokeshi and

Arakaki, 2012) in spite of the potential of habitat complexity to alter food webs, energy

flux, and resilience and resistance of ecosystems to perturbations (Kovalenko et al.,

2012; Floater, 2001). On the contrary, the effects of resource availability on ecosystem

processes is well known, such as on decomposition, food webs, and primary

productivity (Davis et al., 2000; Moore et al., 2004; Fretwell, 1987).

Habitat complexity and resource availability could determine the biological traits of

organisms in a community; therefore, altering ecosystem processes. Biological traits are

organism characteristics at a morphological, physiological, biochemical, phenological,

(37)

37 (Violle et al., 2007) and include body size, locomotion, reproduction, dispersion, and

trophic group (Heino, 2005). Organisms and biological communities influence

ecosystem processes though functional traits (Bello et al., 2010). For example, leaf

traits such as specific leaf area, tensile strength, leaf chemistry, and macroinvertebrate

traits such as body size of detritivores are related with decomposition process (Birouste,

et al., 2012; Gurvich et al., 2003); aerial respiration of invertebrates is related with

oxygen depletion (Dolédec et al., 2006); body size, dispersal ability, and mobility of

invertebrates are related with pest regulation; body size of soil invertebrates is related

with soil stability and fertility (Bello et al., 2010); and, finally, body size is related with

structure and dynamics of food webs (Woodward et al., 2005).

Bromeliad and their fauna can be relevant to assess the effects of habitat complexity and

energy availability on communities and their biological traits, since bromeliads show

variation in both complexity and energy inputs. In terms of complexity, bromeliads

have interlocking leaves where litter and rainfall are accumulated. Moreover, algae,

bacteria, protists, and arthropods live in bromeliad tanks (Benzing, 1990; Brouard et al.,

2012), creating a food web based on detritus and algae (Brouard et al., 2011). Habitat

complexity in bromeliads is measured through leaf number that is related with the

available tank or well number and size of the bromeliad; leaf number joint two aspect of

the habitat complexity the quantitative and qualitative (Stoners and Lewis, 1985). The

amount of litter of allochthonous origin is the principal detrital energy input in

(38)

38 In this paper, we explored the relation between habitat complexity and resource

availability and the biological traits of the invertebrate community in bromeliads. We

tested the hypothesis that habitat complexity and resource availability not only affect

diversity, but also the functional traits of invertebrates such as stage, functional groups,

and dispersal ability. A complex habitat and high resource availability is predicted to

support more species and affect the diversity of biological traits. Moreover, we

predicted differences in abundance and richness between categories of biological traits,

due to adaptations of invertebrates related to colonizing this microecosystem.

2.3. METHODS

2.3.1. Study Area

The study was conducted at the Reserva Forestal Protectora of the Rio Blanco and

Quebrada Olivares Hydrographical Basins. The 4932-ha reserve is located on the western slope of the Central Mountain Range in Colombia (05°3’ 96.7’’ N;

75°26’’88.4’’ W), with a secondary forest and an Alder plantation between 2150 and

3700 m.a.s.l. The maximum annual average temperature is 19°C and the minimum is

6.9°C and the annual mean precipitation is 2500 mm.The most abundant plant families

are Araceae, Actinidaceae, Boraginaceae, and Cyatheaceae.

2.3.2. Methods

We collected individuals of Guzmania multiflora (André) André ex Mez between one

and four meters aboveground. We measured leaf number and the amount of litter

(39)

39 Bromeliad leaves were removed and the water was filtered to collect the inhabiting

invertebrates. The invertebrates were preserved in 70% alcohol and taxonomically

identified to the highest taxonomic level possible using general references: Merrit and

Cummins, 2008; Dominguez and Fernandez, 2009; Stehr, 2005. The biological traits

assessed for the invertebrates were determined by direct observation and the literature.

The biological traits and categories used were: habitat (aquatic, terrestrial), stage (adult,

larvae), functional group (predators, shredders, filter feeders, scrapers, and piercers),

and dispersal type (aerial active, terrestrial active, and aerial passive). Trait abundance

was calculated as the number of individuals per category in each trait, while richness

was calculated as the number of species that had in a given category of a biological trait.

The data was transformed as log (x + 1) for richness and abundance and as log (x) for

leaf number and litter mass.

2.3.3. Statistical analysis

Multiple linear regressions (LM) were used to assess the relation between resource

availability and habitat complexity on taxonomical and biological traits diversity. The

models predicted either richness or abundance of invertebrates associated to bromeliads

or a category of functional trait. We performed a linear regression with all of the

possible combinations of the leaf number and litter amount variables; subsequently, we

selected the best model by the AIC criterion. To compare abundance and richness of the

categories for each biological trait, we conducted a student t-test or ANOVA, if the

traits had more than two categories, by previously verifying assumptions. The

statistical analyses were conducted with the R statistical program.

(40)

40 2.4. RESULTS

We found 4375 macroinvertebrate individuals distributed in 62 morphospecies, of

which 41.93% were aquatic and 61.53% consisted of immature stages. The order

Diptera dominated the community associated to Guzmania multiflora, with the families

Culicidae, Chironomidae, and Syrphidae; however, the most abundant morphospecies

was Scirtes sp. (Coleoptera) with 2439 individuals. Formicidae, Aranea, and Ligidae

were the most frequent terrestrial fauna in Guzmania multiflora (Table 2.1, Figure 2.1).

Abundance and richness differed in relation to habitat complexity; bromeliads with

greater leaf number harbored a higher macroinvertebrate abundance (LM: F2,28= 5.222,

p= 0.011, r2=0.219; log (leaf number) t= 2.344, p= 0.026; log (litter mass) t= 0.762, p=

0.452 and richness (LM: F 2,28= 6.601, p= 0.004, r2=0.271; log (leaf number) t= 3.485,

p= 0.001; log(litter mass) t= -0.843, p= 0.406) (Figure 2.2 A,B).

The categories of biological traits differed in their abundance and/or richness. We

found more abundance of aquatic (t-student; t= -7.084, p= 1.42e-9, df= 63) and

immature invertebrates (t–student; t= -7.094, p= 1.36e-9, df= 63) associated to

Guzmania multiflora, but not more richness of these (richness of adult-larvae t = 0.108,

p= 0.913, df = 63; richness of aquatic-terrestrial t = -0.593, p= 0.555, df = 63).

Regarding the functional groups, the shredders’ group, had mainly by Scirtes sp. and

Chironomidae, showed high abundance in regards to other functional groups, and

richness was high for shredders and predators (ANOVA; abundance: F4, 315= 76.9, p

<2e-16; ANOVA richness F4, 315= 158.4, p <2e-16 (Figure 2.3). Abundance and richness

differed between dispersal types (ANOVA abundance: F2, 186= 64, p <2e-16; ANOVA

richness: F4, 315 = 87.67, p <2e-16, as a greater richness and abundance was found for

aerial active dispersal (Figure 2.3). In addition, habitat complexity affects functional

(41)

41 and predator individuals and aerial active dispersion, as well as a greater richness of

terrestrial, adult, and shredder individuals and aerial and terrestrial active dispersion. On

the contrary, the resource (litter amount) present in bromeliads is negatively related to

richness of filter feeders and terrestrial active dispersion (Table 2.2).

2.5. DISCUSION

Habitat complexity and resource availability determines community structure in

bromeliads (Jocque and Field, 2014; Marino et al., 2013; Srivastava, 2006); moreover,

these factors could determine biological traits. Our study sought to determine the

relation between habitat complexity and resource availability and biological traits of

invertebrates associated to Guzmania multiflora. We predicted that bromeliads with

high leaf number and litter amount supported more invertebrate species and abundance,

as well as more abundance and richness of a particular category of biological traits. Our

results showed that habitat complexity not only alters the taxonomical diversity of

invertebrates in bromeliads, but also the functional diversity of invertebrates through

changes in the abundance and richness of biological traits.

2.5.1. Taxonomical diversity

As we expected, arthropods were the most diverse and abundant group in the bromeliad

Guzmania multiflora, with a high abundance of immature stages of Diptera and high

diversity of functional groups such as shredders, filter feeders, and predators (Merrit and

Cummins, 2008; Stehr, 2005). Immature stages of Diptera have special adaptations to

(42)

42 oxygen from the air and support the low oxygen concentration. Bibionidae (Mestre et

al., 2001), Anisopodidae (Kitching, 1971), Ceratopogonidae (Atrichopogon sp. and

Stilobezzia sp.), Pychodidae (Pericoma sp.) (Campos et al., 2011; Frank et al., 2004),

Chironomidae (Cranston, 2007; Frank, and Lounibos, 2009), and Culicidae have been

reported in bromeliads and/or treeholes. The latter were the most abundant Diptera

families in Guzmania multiflora. Chironomidae is common in bromeliads, where they

have a high abundance and are relevant in energy flow through the ecosystem (Tokeshi,

1995). Culicidae have been well studied given they are vectors of tropical diseases such

as malaria and yellow fever, and most are filter feeders and browse on small particles,

such as Culex sp. and Wyeomyia sp. (Porter and Wolff, 2004). In Guzmania multiflora,

we found the predator Toxorhynchites sp., which is a common predator in treeholes

(Fincke, 1999; Lounibos et al., 2001) and in the bromeliad Aechmea mertensii in French

Guiana (Céréghino et al., 2011).

On the other hand, some morphospecies are a novel record for bromeliads; for instance,

Dixidae individuals feed on small particles in the water column and little is known

about these species in tropical regions (Wagner et al, 2008). In addition, Systelloderes

sp. (Hemiptera: Enicocephalidae) is a predator that lives on the ground or in places

where organic matter in decomposition is accumulated (Schuh and Slater, 1995), and

has been reported in Vriesea inflata and Tillandsia spp. (Bromeliaceae) of Brazil and

Peru (Mestre et al., 2001; Parker et al., 2012). Finally, Oreiallagma oreas (Odonata:

Coenagrionidae) is the only species of Oreiallagma for which its adult stage was

reported in Colombia (department of Valle del Cauca) at 2300 m.a.s.l. in 1918 (von

Ellenrieder and Garrison, 2008), and the biology and ecology of adults and larvae are

(43)

43 Habitat complexity increases richness and abundance of invertebrates in Guzmania

multiflora, given that bromeliads with high leaf number are bigger than bromeliads with

low leaf number. As a consequence, bigger bromeliads have more habitats available to

biota, reducing interspecific and intraspecific competition through providing more niche

space (Cérénghino et al., 2012; Singer et al., 2010), more tanks for retaining more water

and reducing the susceptibility to drought, more canopy litter that is divided among

tanks offering more habitat heterogeneity, and, finally, more adults using the bromeliad

for oviposition (Sota et al., 1994). Litter is a food resource for phytotelmata biota; its

quantity and quality are factors that control population dynamics as well as richness and

abundance of invertebrates (Daugherty and Juliano, 2001). Nevertheless, in our study,

leaf number, rather than litter amount, was a relevant factor on the community, because

fauna chose the tank according to the space available for survivorship and growth of

larvae.

2.5.2. Biological traits diversity

The persistence of species in an ecosystem depends on appropriate morphological,

physiological, and behavioral characteristics that result from species adaptations to

environmental conditions imposed in the ecosystem (Reiss et al., 2009; Paradise, 1998).

For bromeliads, the majority of fauna develops only a part of their life cycle in the

bromeliad; therefore, the immature stage of insects live in the bromeliad by feeding,

moulting, and interacting with other species. Although we found high abundance and

richness of aquatic larvae belonging to the immature stage of Diptera families and

Scirtidae (Scirtes sp.), these were unaffected by habitat complexity and resource

(44)

44 bromeliads or discontinuity in water inputs could determine the diversity of aquatic and

immature fauna.

Nevertheless, richness and abundance of terrestrial invertebrates and adult richness was

altered by leaf number, since the oldest leaves of the bromeliad offer semiaquatic and

terrestrial habitats for terrestrial organisms, which use the bromeliad to consume insects

that arrive to colonize, insects that emerge, and detritus. For instance, Dermaptera and

other terrestrial shredder benefit from leaf litter accumulated in dry leaves; for example,

spiders build webs over the bromeliad or wait at the underside of leaves to predate

insects (obs. per.).

Predators had high richness in the bromeliad and their abundance was related with

habitat complexity, as has been found in other studies (Burlakova et al., 2011;

Langellotto and Denno, 2004). Habitat complexity alters the predator-prey relationship

because habitat complexity reduces predation risk (Saha et al., 2009). However, the

interaction between predators is favored in more complex habitats by increasing

predation efficiency (Grabowski et al., 2008).

Moreover, shredders facilitate other organisms inhabiting the bromeliad through

breaking leaf litter; for instance, shredders regulate the FPOM (fine particule organic

matter) concentration for filter feeders (Heard and Richardson, 1995; Paradise and

Kuhn, 1999;). Although these showed high abundance in bromeliads, shredder

abundance was unaffected by habitat complexity and resource availability, similar to

that found by Paradise (2004), with the most abundant shredder family Scirtidae

(45)

45 to leaf number; more leaves can accumulate more abundance and diversity of food

resources (litter) that can be used by shredders, reducing the predation risk by

damselflies and other aquatic predators, and reducing competition for food resources

and space(McCann and Rooney, 2009).

Other studies have found an increase in the abundance of other functional groups such

as collectors and scrapers in more complex habitats (Burlakova et al., 2011); however,

in our study, scrapers and filter feeder were unaffected by leaf number. On the contrary,

filter feeders richness was reduced where the bromeliad has more litter; due to the litter

decomposition process alter the water pH, which would regulated the presence of filter

feeder, for instance, high litter amount stimulate the detritivores decomposition

(fragmentation) activity over leaves that lead to facilities the presence of Culex sp. and

Wyeomyia sp. through change in the pH (Paradise, 2000; Torreias et al., 2010);

moreover, litter decomposition increase the organic matter available in the bromeliad,

which is essential resource for filter feeder (Kitching, 2001).

Invertebrate with aerial active dispersion are more abundance give that animals with this

dispersion can select and colonizer a new habitat more efficiently those others.

Moreover, the abundance and richness of insects with aerial active dispersion is affected

by habitat complexity, due to insects with aereal active dispersion select the habitat

where ovopositar according to space available for development of immature stage

(Yanoviak , 1998); then insects with aereal active dispersion would select bromeliads

with more leaf number that would receive more detritus and rainwater as a resource for

the survival and develop of the larvae stage (Gename and Monge-Nájera, 2012).

Referencias

Documento similar

The draft amendments do not operate any more a distinction between different states of emergency; they repeal articles 120, 121and 122 and make it possible for the President to

H I is the incident wave height, T z is the mean wave period, Ir is the Iribarren number or surf similarity parameter, h is the water depth at the toe of the structure, Ru is the

In the “big picture” perspective of the recent years that we have described in Brazil, Spain, Portugal and Puerto Rico there are some similarities and important differences,

Since such powers frequently exist outside the institutional framework, and/or exercise their influence through channels exempt (or simply out of reach) from any political

In the previous sections we have shown how astronomical alignments and solar hierophanies – with a common interest in the solstices − were substantiated in the

Percentage of the serve according to the destination area in laterality depending on the competitive level of the couples.. As shown in Table 5, the statistical analysis allows

A hydrogen bond interaction between the o-hydroxy group and the nitrogen of the imine, both present in the dipole, allows the reaction to proceed in the presence of only one

Next, in sections 7 and 8, the influence of the different elements/characteristics of the standard (termination, airline, thermal bead, level of liquid Nitrogen, etc.)